Nafion Interface Cast

Nov 9, 2010 - Cyclic voltammograms of bare Pt(111) electrode in Ar-saturated 0.1 M HClO4 solution and .... Because of side chains are terminated by âˆ...
0 downloads 0 Views 3MB Size
20130

J. Phys. Chem. C 2010, 114, 20130–20140

Spectroelectrochemical Studies of the Pt(111)/Nafion Interface Cast Electrode Ana Ma. Gómez-Marín,*,† Antonio Berná,‡ and Juan M. Feliu‡ UniVersidad Nacional de Colombia, sede Medellı´n, Colombia, and Instituto UniVersitario de Electroquı´mica, UniVersidad de Alicante, Spain ReceiVed: August 12, 2010; ReVised Manuscript ReceiVed: October 8, 2010

Understanding the structure and molecular processes at the electrode/membrane interfaces constitutes an important topic in PEFC as well as in electrochemistry. In this work, the Pt(111)/Nafion model interface in HClO4 acid solutions is studied by IRRAS and cyclic voltammetry. It was found that the presence of an electric field mainly promotes deprotonation of sulfonic groups and structuring of water inside the membrane (polar molecules), especially near the electrode surface, with a sudden change of system optical properties at the Pt(111)/membrane interface at 0.9 V, possibly due to clustering within the polymer. Furthermore, the performance of the Pt(111)/Nafion in a typical electrochemical reaction as CO oxidation has been also analyzed. It is shown that there are notable differences between the characteristics of CO adsorption and oxidation at Pt(111) with and without polymer electrolyte membrane, like a continual wavenumber increase with the potential for the on-top CO band, even during CO oxidation, which proceeds at higher potentials at the electrode covered by the polymer. The spectroscopic features suggest enhanced proton mobility inside the membrane concomitantly with the deprotonation of sulfonic groups near the electrode surface and with higher potentials, possibly due to oriented morphologies inside the membrane induced by high fields. Introduction Polymer electrolyte fuel cells (PEFC) have attracted much interest as one of the most promising technologies for an efficient, nonpolluting power source for vehicles in urban environments. The key element of PEFC is a solid (ionomeric) polymer electrolyte (SPE), which serves as proton conductor and as separator of the anodic and cathodic compartments in this cell. The membrane commonly employed in the most recent PEFC developments is based on Nafion, a DuPont registered trademark polymer. As electrocatalyst, platinum is mainly utilized due to its high electrocatalytic activity in acidic electrolyte environments. In consequence, understanding the structure and molecular processes at the electrode/membrane interfaces constitutes an important topic in PEFC as well as in electrochemistry.1 From a chemical point of view, Nafion, a perfluorosulfonic acid (PFSA) membrane, consists of a hydrophobic poly tetrafluoroethylene (PTFE) backbone with fully perfluorinated ether side chains terminated by strongly hydrophilic -SO3H groups (Scheme 1). This structure leads to spontaneous phase segregation at the nanostructural level. In fact, Nafion can be depicted as a porous solid filled by water.2 In hydrated membranes, the water content is usually stated in terms of the parameter λ, defined as the number of water molecules per sulfonic acid group. It is well accepted that in hydrated PFSA membranes sulfonic acid groups and water develop an interconnected proton conducting network when a threshold value of water content is reached (λ ≈ 2), while the fluorocarbon backbone forms a semicrystalline hydrophobic phase.3 The performance of SPE depends on the coordination of ionexchange groups with water molecules. This coordination results in the formation of ionic channels, through which the mobility * Corresponding author. E-mail: [email protected]. † Universidad Nacional de Colombia. ‡ Universidad de Alicante.

SCHEME 1: Chemical Structure of PFSA 1100 (x ) 7 and y ) z ) 1)a

a

N indicates the nonpolar monomeric units, while P indicates the polar monomeric units.3

of negatively charged ions is dramatically decreased and the transport of positively charged ions (protons) is selectively allowed. Additionally, SPE slows the rate of diffusion for uncharged species, such as methanol and CO2 molecules. This phenomenon plays an important role in direct alcohol fuel cell (DAFC) minimizing the so-called crossover effect.4 Because the membrane transport properties depend on temperature and hydration level,5 the interaction between water and Nafion has been extensively investigated with the help of Fourier transform infrared spectroscopy (FTIRS).6–17 In contrast to the extensive studies of the bulk membrane properties, very little is known about the interaction of the membrane with Pt catalysts, an important question for the understanding of the performance of PEFC.4 It has been recognized that due to the cellular Nafion structure, electrochemical processes can take place only at domains where hydrophilic ionic clusters are in contact with the electrode surface.18,19 The regions where the Pt surface is exposed to the fluorocarbon backbone are inactive. Consequently, when SPE is in contact with a solid electrode, the electrochemically active area is smaller than that of the same electrode exposed to a

10.1021/jp107641r  2010 American Chemical Society Published on Web 11/09/2010

Pt(111)/Nafion Interface Cast Electrode liquid electrolyte. In addition, although the modification of platinum electrochemistry caused by the presence of the Nafion interface is obvious in the voltammetry shown in a number of papers, this aspect has not received significant attention.19–21 It has been reported that Nafion coating significantly enhances methanol electro-oxidation reaction,22 hydrogen anodic oxidation,18 and oxygen reduction reaction -ORR-.23–26 In the latter cases, the increase has been attributed to higher O2 and H2 solubilities in recast Nafion than those in aqueous solutions, which is mainly due to the presence of the hydrophobic fluorocarbon phase.24–26 However, the detailed mechanism of these modifications is still under discussion, and another possible explanation could be a change of the catalyst properties by the membrane-electrode interaction. On the other hand, other studies have concluded that the presence of Nafion does not modify the kinetic parameters of ORR,27,28 or by contrary it has been found that the Nafion-coated Pt system showed lower apparent ORR activity and more H2O2 production than the bare Pt electrode system, indicating that sulfonate groups in Nafion strongly adsorb on Pt sites and modify the surface properties.29,30 These examples illustrate the need of gaining molecular level information related to the nature of the coordination of a Nafion membrane to an electrode surface. In the past, the structure of the Pt polycrystalline/Nafion interface has been probed by different techniques.1,4,31–33 Osawa et al. suggested that the electrochemical microenvironment within Nafion can be investigated by in situ infrared reflection absorption spectroscopy (IRRAS).31 Kanamura et al. investigated it by attenuated total reflection (ATR) configuration, but information about the potential dependent orientation of Nafion chemical groups was not reported because the investigation was conducted in the absence of applied potentials.32 Malevich et al. studied a cast Nafion film on Pt polycrystalline by subtractively normalized interfacial Fourier transform infrared reflection spectroscopy (SNIFTIRS) in the IRRAS configuration, and it was demonstrated that sulfonic acid groups were coordinated to or displaced from the Pt surface depending on the applied potential.4 Later, Malevich et al. investigated the effect of Nafion on CO adsorption and electrooxidation at Pt nanoparticles and observed that a film of Nafion slows the CO oxidation reaction.33 Finally, the interface has been characterized in HClO4 aqueous solutions using surface-enhanced infrared absorption spectroscopy (SEIRAS) by Ayato et al. A potential-dependent band was found around 1100 cm-1 and assigned to the symmetric vibration of the -SO3- groups of the ionomer membrane, and the OH stretching band of non-hydrogen-bonded water molecules associated with the Nafion structure was observed for the first time at 3680 cm-1.1 Single-crystal metal electrodes are often used to characterize electrochemical processes. The rationale behind this is the possibility of establishing a correlation between interfacial properties (i.e., the geometry of the surface, the nature and structure of the adsorbing species) and the electrochemical process occurring on it (i.e., rate of charge transfer, structure of double layer). With this objective, Subbaraman et al. used a voltammetric approach to probe the nature of Pt/Nafion threephase interfaces for Pt(hkl) and polycrystalline platinum surfaces. In that work, they identified, via CO charge displacement measurements, the sulfonate anions as the adsorbing species on the electrode surface.34 Later, Subbaraman et al. studied the kinetics of the ORR at metal/Nafion interfaces on a wide range of surfaces, ranging from Pt(hkl) single-crystal surfaces, Pt-poly, Pt-skin, to high-surface-area nanostructured thin-film (NSTF) catalysts, and showed that the adsorption of sulfonate anions

J. Phys. Chem. C, Vol. 114, No. 47, 2010 20131 on these catalysts negatively impacts the rate of the ORR in perchloric acid solutions but not in sulfuric acid.35 All these approaches provide some valuable information about the possible structure of the interface electrode/membrane. However, great effort still needs to be done to obtain a molecular-level structural description of the three-phase interface. The use of well-defined surface structures together with in situ infrared spectroscopy at electrochemical interfaces provides molecular data on adsorbed species under the influence of the electric field (i.e., the state of internal and external bonds, the lateral interactions within the adlayer). This information gives valuable data for a molecular picture of the electrochemical double layer and contributes to an improved understanding of the physicochemical properties of the electrified interface, which can be correlated with the features in the voltammograms through the potential and coverage dependence of the spectral parameters.36,37 In this work, the Pt(111)/Nafion interface in HClO4 acid solutions is studied by IRRAS and cyclic voltammetry. Special attention has been devoted to the structural changes with potential of the membrane in contact with the platinum surface. Furthermore, the performance of the Pt(111)/Nafion in a typical electrochemical reaction as CO oxidation has been also analyzed. With these techniques, we showed that there are notable differences between the character of CO adsorption and oxidation at Pt(111) with and without polymer electrolyte membrane and that CO oxidation proceeds at higher potentials at the covered electrode. The oxidation of carbon monoxide adlayers on single crystal platinum electrode surfaces is one of the most intensively studied systems, especially the case of the Pt(111) surface. The Pt(111)-CO system can be considered as a model system both in electrochemistry and in the solid-gas interface. The interest in the system arises both from the fundamental point of view of basic science, and because CO is the stable residue formed in the incomplete oxidation of fuels. In the case of electrochemical systems, it is easily formed as the product of dissociative adsorption of simple oxygenated molecules such as formic acid, methanol, ethylene glycol, etc.38 Experimental Section Pt(111) electrodes were prepared following the Clavilier method.39 In all experiments, the electrodes were annealed and subsequently cooled in H2 + Ar gas atmosphere (99.999% -N50Air Liquid in all gases used), protected with a droplet of ultrapure water (Purelab Ultra, Elga - Vivendi), and then transferred to a conventional two-compartment cell. Platinum coil wire and reversible hydrogen electrode (RHE) were used as counter and reference electrodes, respectively. Aqueous solutions were prepared from perchloric acid (Merck suprapur), used as received without further purification (0.01, 0.1, and 1 M). The electrolyte was purged with argon to remove oxygen, and a blanket of argon was kept over the solution during the experiments. Electrodes were assembled into a hanging meniscus rotating electrode (HMRE) holder. Experiments were carried out at room temperature, 22 °C. The time stability of the voltammetric profiles was carefully checked. For SPE deposition, a 5% (w/w) in mixture of lower aliphatic alcohols and water solution of Nafion (Aldrich) was diluted with ultrapure water to a concentration of 1% (w/w). The Nafion solution was dropped and spread over the Pt(111) surface uniformly. The membrane was deposited by a spin-coating procedure at 750 rpm in an Ar atmosphere. Typical drying times ranged from 15 to 30 min. We did not measure the film

20132

J. Phys. Chem. C, Vol. 114, No. 47, 2010

Figure 1. Cyclic voltammograms of bare Pt(111) electrode in Arsaturated 0.1 M HClO4 solution and Nafion-covered Pt(111) electrodes in Ar-saturated 1, 0.1, and 0.01 M HClO4 solutions at a scan rate of 50 mV s-1. Insets: Detailed views of the new pseudocapacitative processes at 0.45-0.55 V at different pH’s.

thickness, but it can be calculated from the concentration and the volume of Nafion solution deposited on the electrode and by using a dry density of 1.5 g cm-3. Typical variations in thicknesses correspond to about 0.88-1.7 µm (they were not measured and were calculated by using an apparent density for the Nafion film). The Nafion-coated electrode was then cured at 135 ( 5 °C for 1 h to evaporate the solvent and to improve the adhesion between the polymer film and the electrode.1,4,33 Infrared spectroscopic experiments were performed in a Nicolet Magna 850 infrared spectrometer following the same experimental protocol used before.40 p- and s-polarized lights were used in these experiments. For CO adsorption experiments, only p-polarized light was used. As in electrochemical experiments, the solution was purged with argon to remove oxygen, and a platinum coil wire and a reversible hydrogen electrode (RHE) were used as counter and reference electrodes, respectively. A BaF2 infrared window was used to reach vibrational wavenumbers down to 900 cm-1. The spectra were shown as the ratio -log(RES/RER), where RES and RER are the single beam spectra obtained at the sample and the reference potentials, respectively. Carbon monoxide was adsorbed at 0.1 V by bubbling CO (99.997% -N47-) gas through the solution. The excess of dissolved CO was removed by purging with Ar. The characteristic hydrogen adsorption/desorption region was completely blocked by the adsorption of CO. The reference potentials were set as +100 mV for Pt(111)/ Nafion experiments and +900 mV for CO-Pt(111)/Nafion experiments (CO was completely oxidized at that potential) and +100 mV for CO2 (this potential is more negative than the onset potential of the oxidation of CO to CO2). Results and Discussion Pt(111)/Nafion Interface. Cyclic Voltammetry. Pt(111)/ Nafion interface was initially characterized with cyclic voltammetry. Nafion film showed good stability and electrode adherence. The first CV recorded after inserting the electrode into the cell showed that Pt(111) surface was blocked. Figure 1 shows the CVs of Nafion-free (0.1 M HClO4) and Nafioncovered Pt(111) (1, 0.1, and 0.01 M HClO4) in different HClO4

Gómez-Marín et al. solutions after applying multiple potential cycles between 0.06-0.90 and 0.06-1.2 V for several minutes (2-4 min), as cleaning procedure. Up to 0.90 V, the electrode surface reaches a stationary condition, while a more positive upper potential limit slowly increases the active surface area, indicating a modification in the membrane distribution near the electrode surface due to changes in the metal-ionomer interactions. After repetitive cycling up to 1.2 V, surface defects grow on Nafionfree and Nafion-covered Pt(111) electrodes. In the next results presented in this work, only potential cycles between 0.06 and 0.90 V were applied to the covered electrode as a cleaning procedure. Actually, the state of the art of the different adsorption states of a well-ordered Pt(111) electrode in acidic solutions without anion specific adsorption, as perchloric or trifluoromethane sulfonic acid solutions,41 recognizes two zones that can be clearly distinguished (Figure 1). At potentials below 0.40 V, pseudocapacitative currents originated by hydrogen adsorption/ desorption processes (denoted as underpotential deposition of hydrogen, H+ + e- f HUPD) are observed. At potential above 0.55 V, the pseudocapacitative current observed has been traditionally assigned to the water dissociative adsorption as OH species.42 In the case of Nafion-coated Pt(111) (Figure 1), these two zones are still differentiated, but they are modified quantitatively, especially at lower pH solutions: the hydrogen and OH adsorption/desorption regions are significantly blocked, probably because a fraction of the Pt(111) surface is covered by hydrophobic PTFE backbone and becomes electrochemically inactive.1,4,18 Careful inspection of the CVs reveals a significant feature on the Nafion-covered surfaces: a small, new pseudocapacitive process, relatively irreversible and pH dependent in the potential region between 0.45 and 0.55 V. This behavior could indicate specific interaction between the membrane and electrode surface, markedly visible at pH ) 2.0. This result is in contrast with the common consideration of Nafion as a nonadsorbing electrolyte accepted in the literature,18,20 but agrees with recent works of Subbaraman et al. on Pt(111)34,35 and other works with polycrystalline Pt, which suggested strong adsorption of sulfonate groups in Nafion on Pt sites.29,30,33 Apparently, the reversibility, peak position, and the extent of the adsorption of the new charge transfer process are strongly pH dependent; however, these results should be taken with caution: although the pH within the Nafion coating is not altered by changes in the pH of the solution, the measured peak potential can change in response to the changes in the ionic strength of the solution, which affect the Donnan potential43,44 and Nafion acidity.45 On the other hand, in the absence of Donnan potential effects (i.e., high ionic strength of supporting electrolyte), process irreversibility could be due to different causes: the reduced mobility and the limited spacing between the anions arising from the structure of the ionomer membrane, similar to the suggested by Subbaraman et al.,34 or the development of an estable membrane morphology, induced by the electric field, which needs higher perturbations to come back to the “original” state.46,47 In the same form, pH near the electrode surface should be less than the pH of the solution, considering Nafion acidity,45,48–50 and CVs presented in Figure 1 are slightly shifted to lower potential values. The extent of this shift also depends on pH and ionic strength of the electrolyte, which modifies cluster size, water structure, and ion solvation inside the membrane.43–45 Alternatively, higher proton discharge currents during the cathodic swept could suggest an enhancement of this reaction

Pt(111)/Nafion Interface Cast Electrode

J. Phys. Chem. C, Vol. 114, No. 47, 2010 20133

Figure 2. Cyclic voltammograms for a Nafion-coated Pt(111) in Arsaturated 0.01 M HClO4 solution at a scan rate of 50 mV s-1 and different upper and lower cyclic voltammetry limits. Black and magenta, 0.06-0.90 V; red, 0.06-0.70 V; and blue, 0.30-0.90 V.

Figure 3. Cyclic voltammograms for a Nafion-coated Pt(111) electrode in 0.01 M HClO4 + xM HSO4- (x ) 0, 10-2, and 10-1 M) at 50 mV s-1. For the sake of comparative purposes, in the inset is included the corresponding CVs for a bare Pt(111) electrode.

at covered electrodes; however, much more work is necessary to clarify this point. Experimentalists are familiar with the fact that the final structure of an ionomer membrane is dependent on the method by which it is prepared, and that the final form of the material is rarely the configuration with the lowest free energy. Boiling, rinsing, microwave drying, and other thermomechanical steps, as well as the manner of removing contaminants from the membrane, influence the membrane structure.46,47,51 The results in Figure 1 show marked contrast with the findings of Subbaraman et al., who reported no blockage of hydrogen adsorption/desorption regions and measured bigger charge transfer for the pseudocapacitive process.34 Thickness and pretreatment differences could account for these discrepancies. Additional experimental work about the electrochemical properties of this charge transfer process also showed that the electrochemical reversibility, peak potential, and height depend on the value of the lower potential limit of the CV (Figure 2) and the thickness of the film (coverage), as also was suggested by Subbaraman et al.35 The larger voltammetry peaks with thicker films are probably due to the larger amounts of redoxactive ions localized in the film even though the concentration is the same as in thin films. This is consistent with the behavior of reversible couples in thin layer cells in which the peak current is directly proportional to volume. On the other hand, the upper limit of potential of the CV and the scan rate do not affect this process (Figure 2). Attending to Nafion chemical nature and that the peak potential is located in the same potential region where the adsorption of (bi)sulfate anions is observed in sulphuric acid solutions on Pt(111), the charge transfer process taking place between the membrane and the metal can be tentatively assigned to the oxidative adsorption of sulfonate anion side chains, as it has been suggested in previous works:4,34,35

protons, and metal electrode, there are many metastable states separated by energy barriers in the equilibrated membrane, and the questions of how they affect the membrane morphology and the charge transfer at the metal/ionomer interface are not yet answered. Although a few works have provided some valuable information,1,4,18–35 a great effort still needs to be done to obtain a molecular-level structural description of the three-phase interface and develop more detailed and complex models than the simple “spring model” proposed by Subbaraman et al.34,35 Because of side chains are terminated by -SO3H group, Nafion dramatically decreases the mobility of negatively charged ions through the membrane and eventually can suppress it. To gain further insight into the anion permeability, the voltammetric modifications caused by the addition of sulfate anions at 10-3, 10-2, and 10-1 M in a perchloric acid solution (0.01 M) have been recorded, Figure 3. Sulphuric acid was used as additive because the adsorption of sulfate anions on Pt(111) has a characteristic voltammetric response. Figure 3 shows that at low sulfate concentrations (∼10-3 M) the membrane suppresses anion permeability, while at higher concentrations the sulfate anions are able to reach the electrode surface. The exact concentration of anions diffusing into the membrane depends on film thickness and the ionic strength of the solution, which determines the Donan potential at the interface between ionomer and the solution.43,44 Apparently, sulfate adsorption suppresses Nafion sulfonate adsorption. It appears that at 0.1 M the ionic strength of the solution is enough to compensate the Donnan potential effects, with the common thickness employed in this work. IR Band Assignment. The assignment of the vibrational bands of Nafion was based on the already published data.6–17 The fundamental vibrational modes of Nafion appear below ∼1500 cm-1, and the frequencies of the vibrational modes corresponding to the -SO3- group symmetry (i.e., on its environment) are expected in the 1100-1400 cm-1 range. Unfortunately, this IR wavenumber range is obscured by the intense CF2 stretching vibrations, and therefore it is difficult to make a precise assignment.9,12,13,17 Warren and McQuillan noted the importance of considering vibrational contributions from more than one functional group when assigning IR absorptions of fluoropolymers.17 The consideration of mechanically coupled internal coordinates is

-SO3H f -SO3Pt + H+ + 1e-

(1)

Analogous to (bi)sulfate adsorption/desorption on Pt(111), the peak potential position depends on SO3- anionic concentration and pH of the solution. Nevertheless, due to the complexity of interactions between the membrane monomers, water molecules,

20134

J. Phys. Chem. C, Vol. 114, No. 47, 2010

TABLE 1: Selected Infrared Absorption Bands of H-Nafion6–17 band location (cm-1) 970 m 980 s 1060 s ∼1160 s 1200-1225 vs,a brb 1130-1300 s 1300 sh 1320 sh 1430

assignment C-O-C symmetric stretching, mechanically coupled to -HSO3: νs(C-O-C), which is sensitive to membrane hydration νs(C-O-C) ether moiety close to the polymer backbone symmetric stretching of -SO3- groups, mechanically coupled to C-O-C: νs(SO3); hydrogen bonded to H2O molecules asymmetric stretching of -CF2 groups νas(CF2) mixed region CF/SO3-. CF2 stretching, symmetric: νs(CF2); antisymmetric stretching modes of -CF2, νas(CF2) νas(SO3-) νas(CF3) stretching mode νs (SdO) of -SO3-15 stretching mode νas (SdO) of -SO3H15

a Relative intensity: vs, very strong; s, strong; m, medium; w, weak; sh, shoulder. b Band width: br, broad.

essential for the analysis of infrared spectra of ionomers and correlation of those spectra with the effects of ion exchange and state of hydration.16 The main absorption bands and corresponding assignment are listed in Table 1. It should be recognized here and throughout the discussion that in a system such as that studied, there is much mixing of the vibrational modes due to the similar masses and symmetry of the various components of the molecule, especially the COC, CF2, and C-S groups. This mixing complicates spectral assignment based on an interpretive model centered around isolated local modes. Thus, the listed assignments are the dominant vibrational mode(s) that contribute to any given absorption.6–17 The region above 1500 cm-1 is free of bands from the strong fundamental vibrational modes of the polymer. The rest of the bands in the region 4000-1500 cm-1 enable the observation of features attributed to permeating molecules (i.e., water vapor and other types of solvents), mainly to condensed water in the membrane.11,12 It has been proposed that three types of water can exist: not hydrogen bonded, ∼3715 cm-1; partially hydrogen bonded, ∼3660 cm-1; and hydrogen bonded (either to another molecule of water or to a sulfonate group), ∼3520 cm-1 (Table 2).7 In Situ IR Reflection Spectra. Once the electrode was immersed in 0.1 M HClO4 solution, p-polarized light spectra were recorded to determine the changes during membrane hydration processes. Figure 4 shows IRRAS spectra of Nafioncovered Pt(111) in deoxygenated 0.1 M HClO4 solution at open circuit potential, as a function of the time elapsed since the electrode was put in contact with solution. With increasing hydration, different types of water inside the membrane are observed: a bipolar band at 3068-3360 cm-1 suggests a wavenumber shift progressing into more bulklike water OH stretching (3400 cm-1). The intensity of this band increases with time, following the increase and sharpening of the band at 1650 cm-1. This behavior can be associated with the phase separation due to membrane swelling because of membrane hydration (increase of the hydrophilic region). In addition, the band intensity increase at 3880 and 3630 cm-1 implies an increase of population of “free” and partially hydrogen-bonded water due to the water intercalation into the polymer backbone and the subsequent growth in the interfacial region. Alike movements have been obtained from different polymer hydration studies,11,12 and the experimental timedependent absorbance was fit by a pore diffusion model, consistent with proposed structures of Nafion as consisting of a network of hydrophilic pores and channels.12

Gómez-Marín et al. TABLE 2: Selected Infrared Absorption Bands of Water inside H-Nafion6–17 band location (cm-1) ∼1625-1635

∼1735

∼3080 ∼3520

∼3660 ∼3715

assignment close in energy to the bending mode of bulk water (1645 cm-1) and is typically observed somewhat below in IR spectra of hydrated, metal cation exchanged Nafion materials characteristic of solvated proton structures, (H2O)nH+; it has been assigned to asymmetric H-O-H bending modes of solvated H3O+ species H2O strong bound (icelike structure) water that forms hydrogen bonds, either to another molecule of water or to a sulfonate group from Nafion, the strength of which are significantly weaker than that of the hydrogen bonds in pure water partially hydrogen bonded: water that is partially bounded to fluorocarbon and has one proton available for hydrogen bonding non-hydrogen-bonded water: water with both protons surrounded by fluorocarbons not forming any hydrogen bonds, oriented in a manner that extends both hydrogen atoms toward fluorocarbon regions of the polymer or possibly at the interface of water and air-filled voids in pores and channels of the polymer

Concomitant to water uptake, the increase of the bands at 986, with the shoulder at 970 and 1170 cm-1, and the 1200-1300 cm-1 region, directly associated with the polymer backbone (Table 1), points toward global membrane reorganization. Similar trends have been reported from time-resolved spectral measurements for membrane hydration with Na+exchanged Nafion.12 The rapid uptake of water precluded similar studies into H+-exchanged Nafion.12 In the former case, leastsquares modeling was applied to gain insight into changes that occur in the structure of polymer membrane during hydration with an excellent match for bands of the CF2 and C-O-C groups modes, but agreement was not as close for bands arising from modes of the hydrophilic -SO3- and water.13 No further efforts were tried to fit a similar model in our case. The growth of bands at 1060 and 1320 cm-1 with time indicates that the sulfonate groups shift to a deprotonation state

Figure 4. IRRAS spectra for a thin Nafion film cast at a Pt(111) electrode at open circuit potential versus hydration time. The reference spectrum was set as corresponding to t ) 0 s. In the inset is included the spectral region up to 3600 cm-1 to show the OH stretching potentialdependent behavior.

Pt(111)/Nafion Interface Cast Electrode

J. Phys. Chem. C, Vol. 114, No. 47, 2010 20135

Figure 5. IRRAS spectra for a thin Nafion film cast at a Pt(111) electrode. The reference spectrum was taken at 0.1 V (RHE). Potential sweep rate 2 mV s-1. Left: p-polarized light. Right: s-polarized light. In the inset is included the spectral region up to 3600 cm-1 to show the OH stretching potential-dependent behavior.

-SO3- and are mainly dissociated at the equilibrium. Conversely to the reported, the intensity of these bands is bigger than similar bands from the spectral response either during polymer hydration studies without the electrode surface11–13 or for a Nafion film cast at the Pt polycrystalline electrode,4 which suggest a strongest interaction between -SO3- ions and the Pt(111) surface. Finally, the band at 1100 cm-1 is assigned to perchlorate ion inside the membrane, as a result of permeation from the supporting electrolyte. With the exception of the band at 3068-3360 cm-1, the wavenumber of the other bands is very little affected by the hydration time. Influence of the Potential on Nafion Membrane. Once full membrane hydration is reached, cyclic voltammetric sweeps were performed between 0.1 and 0.9 V at a scan rate of 2 mV s-1, to study membrane changes with the applied potential at the interface polymer/Pt(111) as well as in the bulk of the membrane. Figure 5 shows the IRRAS spectra for Nafioncovered Pt(111) in 0.1 M HClO4, every 100 mV, taking as reference potential the initial spectrum at 0.1 V. The absence of characteristic bands due to bulk water at 3400 and 1645 cm-1 implies that the signals from the bulk solution were successfully canceled out by the polymer. As can be observed from Figure 5, the presence of an electric field mainly affects the water distribution and sulfonic groups inside the membrane (polar molecules) and promotes the perchlorate anion permeation. In addition, the wavenumber of the bands in the spectra collected either with p-polarized or with s-polarized light is not too much influenced by the potential (e.g., there is no observable Stark effect). The changes in the membrane with the potential are very slow: reversing the potential sweep direction does not affect significantly the trends of the membrane absorption changes. This fact suggests the development of quasi-irreversible metastable states induced by the electrode potential at the interface at room temperature, as was recently proposed from molecular simulations.46,47 As was mentioned before, in systems as complex as humid ionomers, it is not unexpected that there should be several different structures that each represent a local minimum in the free energy and that the barriers to transitions between these morphologies should be much larger than available thermal energies.46,47 Applied perturbations such as the external field in our study or raising the membrane temperature52,53 are able to drive the system to states with lower energies and internal stresses in the backbone membrane.

Absorbance in the vicinity of 2360 cm-1 is believed to arise from overtones of polymer skeletal modes and low wavenumber vibrations associated with O-H groups.12 Above ∼0.7 V during the anodic sweep, a band at 2344 cm-1, related to CO2, starts to be detected, and its intensity follows the potential sweep, decreasing during the negative-going sweep down to ∼0.7 V. The formation of CO2 can be originated as a product of the oxidation of solution impurities from the cast film or, more likely, to incipient membrane degradation with the potential. More work is necessary to clarify this point. In particular, the applied potential has a direct effect on the structure and size of the hydrophilic ion cluster region of the membrane that can be understood following the changes in the water absorption bands, as it has already been reported.31 The number of water molecules playing a role in the hydrogenbonding network inside the membrane, but weaker than in bulk water, increases, while the concentration of isolated (“free”) water (3720 cm-1) decreases with the potential. This suggests a growth in the hydrated ion cluster region in addition to changes that may occur in the environment for water surrounding ionic groups within the polymer and decreased area for water/polymer fluorocarbon region interface and explains the increase in the active surface area during successive CVs up to 1.2 V. On the other hand, the appearance of a sharp band at 3180-3200 cm-1, characteristic of very strong hydrogen-bonded water molecules like those in the ice-network structure,36 and the feature near 1650 cm-1 (at open circuit potential), close in energy to the bending mode of bulk water (1645 cm-1), shifted downward to 1620 cm-1 imply stronger interactions between water molecules and the electrode surface on the Pt(111)/Nafion interface at higher potentials, probably because the hydrogenbond water network inside the membrane is weaker than that in bulk liquid water. The lower energy mode as compared to bulk water has been discussed in terms of water adsorbed on the electrode surface with a electron lone pair54 or weakening in the hydrogen-bonding network caused by interactions of water molecules with sulfonate ions in Nafion.12 Hence, it seems that both stronger water/metal interactions and sulfonate adsorption may lead to the OH adsorption to be shifted to more positive potentials on the CVs recorded with Nafion-covered electrodes. In a similar way, a broad band at 1780 cm-1, generally assigned to H-O-H bending modes of solvated H3O+ species, starts to increase in intensity at more positive potentials, possibly due to oriented structures inside the membrane induced by high external electric fields as has been reported before.46,47 This

20136

J. Phys. Chem. C, Vol. 114, No. 47, 2010

Gómez-Marín et al.

Figure 6. A series of IRRAS spectra recorded for Nafion-coated Pt(111) electrode surface. The reference spectrum was taken at 0.10 V (RHE), and the sample potential E2 varied from 0.10 to 0.90 V. Left: p-polarized light. Right: s-polarized light. In the inset is included the spectral region from 3000 to 4000 cm-1, to show the behavior of the OH stretching band. The rest of the experimental parameters are the same as in Figure 4.

points out enhanced proton mobility inside the membrane concomitantly with the deprotonation of sulfonic groups, that is, an increase in the proton conductivity and membrane acidity, as was also found in previous works on stretched membranes and on Nafion poling experiments.52,53 This phenomenon originates from the various forces acting upon the membrane and could be the origin of major H2O2 formation during the ORR in presence of Nafion, due to that lower pH’s raise H2O2 production, as has previously been reported.23,29,55 To elucidate which are the functional groups directly interacting with the electrode surface, p-polarized and s-polarized light infrared experiments were compared. Strong water movements near the Pt(111)/Nafion interface are testified by the growth in intensity of the band at 3200 cm-1 in the spectra recorded with p-polarized light. In addition, p-polarized light spectra present two growing bands at ∼1230 and 1430 cm-1, which are absent in the s-polarized light spectra, and can be related to -SO3and -HSO3 groups, respectively. In this way, more positive potentials induce an orientation in the membrane sulfonic groups nearest the electrode surface. Likewise, the -HSO3 anion band implies a redox equilibrium on the electrode surface and confirms the sulfonate groups as the adsorbing species on Nafion-coated electrode between 0.45 and 0.55 V, analogous to (bi)sulfate adsorption equilibrium on Pt(111).34–36 Sulfate/bisulfate adsorption on Pt electrode surface has been systematically investigated and characterized by IRRAS technique.36,56 The spectra for bisulfate ions on Pt electrode surface exhibited the blue-shift with increasing electrode potential, while for sulfate adsorption on Pt(111), a weak redshift of the peak 1120 cm-1 was observed with potential increase.1,56 The lack of a wavenumber shift in this study would suggest that sulfonate anions are not specifically adsorbed but populated in the double layer. However, confined polymer environment or an assisted charge transfer by water molecules can obscure this movement. Additionally, it is notable that for Nafion, combined infrared spectroscopy and density functional theory calculations on a model side chain compound revealed that many modes dominated by side chain C-F stretching also contain coupled motion of atoms in the C-O-C and -SO3functional groups,15–17 especially changes within the 1250-1100 cm-1 region, where the asymmetric stretching mode(s) of -SO3are known to absorb.16,17 More work is necessary to clarify this point. The band at 1100 cm-1, assigned to perchlorate ion inside the membrane, increases with the electrode potential, indicating

that ionomer cannot exclusively support the electrical double layer and thus has contribution from bulk electrolyte ions. Anions can penetrate into Nafion structure and probably reach the Pt surface possibly because of the high concentration of the acids employed in the present study (0.1 M HClO4), in a way similar to the case of SO4- anions. Figure 6 shows p-polarized and s-polarized light IRRAS spectra for successive potential jumps between 0.10 and 1.20 V, every 100 mV, for Nafion-covered Pt(111) electrode in 0.1 M HClO4. The reference spectrum was taken at 0.10 V. Between 0.10 and 0.90 V, the spectra are very similar in both Figures 4 and 5, indicating that the membrane is not fully hydrated. In addition to the bands discussed before, s-polarized light spectra have a new band at 1450 cm-1, characteristic of protonated sulfonic groups. This band is not clearly seen in Figures 4 and 5 and indicates a small fraction of protonated terminal side chain groups. It should be noted that there is a sudden increase in the band intensity for almost all bands above 0.9 V in p-polarized light spectra (Figure 6), especially in the bands of water, and this could be related to a change in the optical properties of the metal/membrane interface. This phenomenon can be associated with a change in the membrane/metal interactions at the beginning of the surface oxidation at this potential, similar to the effect discussed before in the cyclic voltammetry experiments. In a previous work, Malevich et al. related the band intensity of a bipolar band with its peak-to-peak amplitude difference and demonstrated its proportionality to the surface concentration of the species responsible of that band.4 The peak-to-peak amplitude difference of the bipolar bands related to OH stretching and bending water modes in the p-polarized and s-polarized light spectra is plotted against electrode potential in Figure 7. Initially, peak-to-peak amplitude grows almost linearly with the potential for both types of polarized light. At 0.9 V, the curve for p-polarized light presents a jump, indicating a sudden increase in band intensity in this region (Figure 7). Osawa et al. have reported experimental evidence of the formation of a stable icelike structured water on positively charged platinum surfaces. This conclusion was based on the appearance of a broad band at ∼3000 cm-1 in the ATR-SEIRAS spectrum in 0.1 M HClO4. It was suggested that this band is expected to be shifted from ∼3000 to ∼3200 cm-1 on an oxidized Pt surface, if water molecules are hydrogen bonded to surface oxide, and would considerably increase in intensity by surface oxidation.

Pt(111)/Nafion Interface Cast Electrode

Figure 7. Peak-to-peak amplitude of bipolar bands in the p-polarized (s-polarized) light spectra. Left: Water bending, 1570-1650 (1565-1640) cm-1. Right: OH stretching, 3100-3355 (3020-3335) cm-1, as a function of electrode potential (data extracted from Figure 6).

However, as in their work, the icelike structure was stable even in the surface oxidation region. They gave as a possible explanation that the Pt atoms always appear on the oxidized surface due to the place exchange between the adsorbed oxygen species and underlying Pt atoms.54 On the basis of the results discussed above, we suggest that the band shift from ∼3100 to ∼3200 cm-1 (Figure 6) together with growth of the peak-to-peak amplitude difference (Figure 7) at Nafion/Pt(111) interface can be a clear evidence of the water icelike structure disruption. This rupture can be originated because the membrane prevents the place exchange of the Pt(111) atoms during surface oxidation, or because this surface structure is weaker than at the Pt(111)/solution interface, due to the weaker hydrogen-bonding network inside the membrane as compared to bulk water. CO-Pt(111)/Nafion Interface. Cyclic Voltammetry. The electro-oxidationofsaturatedPt(111)/Nafion-COandPt(111)-CO adlayers was investigated. The CO adlayers were prepared by exposure of the electrode surface to CO gas dissolved in 0.1 M perchloric acid solutions at 0.10 V in an oxygen-free atmosphere. After the formation of a saturated adlayer of CO, witnessed by the low double layer current recorded in the CV between 0.10 and 0.30 V, Ar was bubbled through the solution for at least 15 min to remove traces of dissolved CO. During the removal of CO from solution, the working electrode was kept at open circuit potential above the solution. Figure 8 shows the CVs recorded at 50 mV s-1. As it can be seen in the inset of Figure 8, when the CO adlayer is fully oxidized, the CVs for both solutions recover the original profile. CO oxidation at the Pt(111)/Nafion interface at 50 mV s-1 takes place at more positive potentials and within a narrower peak as compared to the Pt(111) electrode (Figure 8): ∼0.87 versus ∼0.77 V. This behavior suggests that Nafion influences reaction environment and slows CO oxidation. This could be caused by a structural modification of the CO adlayer or by a change in the adsorption conditions of the oxygenated species necessary for CO oxidation, that is, adsorbed OH,33,57 similar to the CO oxidation in sulphuric acid solutions on Pt(111).58–60 At 20 mV s-1, the CV has two consecutive sharp peaks for the CO oxidation, the fist one at ∼0.82 V and the second one around ∼0.84 V (data are not shown).

J. Phys. Chem. C, Vol. 114, No. 47, 2010 20137

Figure 8. Cyclic voltammograms for saturated Pt(111)/Nafion-CO and Pt(111)-CO adlayers in 0.1 M HClO4 solution. Potential sweep rate 50 mV s-1. Inset: Detailed view of the double layer region to show the adsorption-desorption processes.

During the negative-going sweep, in the CV of the Nafioncovered electrode appears a broad reduction peak, centered at ∼0.36 V, inset Figure 8. This peak could be related to bicarbonate anion desorption that comes from CO2 produced during CO oxidation.37 Bicarbonate adsorbs on the electrode due to a major residence time near the metal surface because of its slower diffusion rate through the membrane; this was also evident from IRRAS measurements (left, Figure 10). Slow CO2 diffusion may be a serious drawback in other Pt surface sites because CO formation would be possible.61–63 Active surface area of Nafion/Pt(111) electrodes was estimated by integration of the hydrogen adsorption and CO oxidation charges from the CVs. The active surface area determined by both methods was always lower for Nafion/ Pt(111) as compared to the Pt(111) interface; that is, the ratio between the charges for CO-Pt(111)-Nafion/CO-Pt(111) was 0.70 at 50 mV s-1 and 0.75 at 20 mV s-1. Furthermore, the active surface area was always lower when determined from hydrogen adsorption charge as compared to that from CO oxidation charge, at least as much as twice lower. This suggests a Nafion layer shift by the adsorbed CO that is confirmed by an increase in the hydrogen adsorption charge in the following potential cycles after CO oxidation. With successive cycling, the membrane slowly recovers its initial distribution near the electrode surface. However, the increase in the charge corresponding to hydrogen adsorption, that is, the increase in active area for this process, is not higher than the active area determined from CO oxidation charge. Hence, it can be concluded that the membrane preferably blocks surface sites with adsorption strength that can be displaced by CO adsorption but not by hydrogen adsorption. In Situ IR Reflection Spectra. In situ IRRAS has been employed to investigate CO adsorption and oxidation at Nafioncoated Pt(111) interface. A voltammetric sweep was performed between 0.1 and 0.9 V at a scan rate of 2 mV s-1. Figure 9 shows IRRAS spectra of CO for free and Nafion-covered Pt(111) electrodes in 0.1 M HClO4 at 0.5 V, obtained by choosing as reference potential the single beam spectrum collected at 0.9 V. There are significant differences between the CO spectra on both electrode surfaces. The most striking characteristic for Nafion/Pt(111) interface is that no bridge-bonded CO band, at

20138

J. Phys. Chem. C, Vol. 114, No. 47, 2010

Figure 9. IRRAS spectra of CO adsorbed on Pt(111) in Ar-saturated 0.1 M HClO4 solution at E ) 0.5 V (RHE). Scan rate of 2 mV s-1. The reference spectrum was taken at 0.9 V (RHE).

around 1840 cm-1, was observed.31,33 Likewise, the wavenumber of linearly bonded CO is blue-shifted for covered electrode, 2075 versus 2067 cm-1, contrary to the results reported by other authors for polycrystalline Pt.33 The CO2 band is similar in both spectra as it would be expected for a nonadsorbed species. Considering that the linear CO band shifts to higher frequencies with decreasing pH for bulk Pt electrodes,59,60,64 and Nafion is a super acid,45,48–50 the band shift on the polymer-covered electrode points toward local pH differences between Pt(111) and Pt(111)/Nafion interfaces, that is, higher proton activity in the nearest surface layer on the membrane covered electrode, in accordance with previously discussed results (increase of the band intensity at 1780 cm-1 with higher potentials). Alternatively, the difference between the wavenumber of the linearly bonded CO for membrane covered and membrane free electrodes could be caused by a different packing density in the CO adlayer, because of lower CO coverage or fewer water molecules around each CO molecule at the membrane hydrophobic regions in the first case as compared to in the second case (wavenumber closer to that of gaseous CO, ∼2100 cm-1).57,65,66 We would like to point out that while CO oxidation on Pt(111) starts at ∼0.45 V, with a maximum CO2 intensity at

Gómez-Marín et al. ∼0.65 V (left, Figure 10), on the Nafion-covered electrode this oxidation process is shifted to ∼0.65 V, with a maximum CO2 intensity at ∼0.85 V (left, Figure 10), similarly to that reported by Malevich et al.33 Above ∼0.60 V, the CO oxidation is completed for the bare Pt(111) electrode, and the band intensity of CO2 quickly decreases with time (left and right, Figure 10). For the Nafioncovered electrode, CO oxidation is completed at ∼0.90 V, and CO2 inside the membrane remains up to potentials below 0.10 V during the cathodic sweep (left and right, Figure 10). Therefore, the CO2 diffusion rate is slower through the membrane than in bulk water and facilitates bicarbonate absorption at approximately 0.36 V, as it can be seen on the CV (Figure 8). The intensity and wavenumber of CO band also depend on the electrode potential (center and right, Figure 10), while the bandwidth for both CO and CO2 has similar behavior at both interfaces during the whole potential sweep. In the case of CO molecules, the full width at half-maximum is around 14-15 cm-1 and is almost independent of the electrode potential and CO coverage, until the beginning of the CO oxidation is reached.59,60,65 Similar behavior has been observed for COstripping on a Pt(111) electrode by Chang and Weaver56 and for a polycrystalline Pt electrode by Malevich et al.33 This could indicate that CO oxidation at the Pt(111)/Nafion interface takes place through island formation, nucleation, and growth kinetics, because the environment of CO molecules inside the islands remains essentially unchanged with coverage.33,56 During the whole potential sweep, the wavenumber for linearly bonded CO on the Nafion-covered electrode is always greater than that on the bare Pt(111) electrode (center, Figure 10) and continuously increases with potential at 29 cm-1 V-1 (Stark tuning rate). This shift is consistent with the previous literature reports for Pt bulk electrodes,31,33 while for Pt(111) the slope is 34 cm-1 V-1 up to the onset of CO oxidation, similar to that reported for a coverage of 0.65 ML approximately.36,54,65 The oxidation of adsorbed CO requires that water molecules penetrate the compact CO adlayer, providing oxygen needed for the conversion of CO to CO2. This reaction is slower for adlayers that display stronger dipolar coupling.33 From the results discussed above, the Stark effect is slightly larger for the bare surface than for the covered electrode, and this could imply that the dipole-dipole coupling effect, operating between the dipoles of the CO molecules adsorbed on the surface covered

Figure 10. CO IR bands analysis for bare Pt(111) (red) and Nafion-covered Pt(111) (black). Left: CO2-integrated band intensity versus the electrode potential. Center: Wavenumber of the CO band at maximum intensity. Right: CO-integrated band intensity versus the electrode potential.

Pt(111)/Nafion Interface Cast Electrode with the ionomer membrane, is stronger than that on the bare Pt(111). This, concomitantly with the presence of less water molecules around the surface CO islands on the membrane hydrophobic region, would cause the upward shift of the wavenumber for linearly bonded CO during CO oxidation. The current theory for CO oxidation takes into account that its rate-determining step is the formation of a hydrogen-bonded water-OHads network, strongly influenced by anions, and that CO oxidation occurs, at least in part, by the diffusion of OHads through this network.67 To understand why a stronger dipolar coupling exists at the Nafion/Pt(111) interface, it can be tentatively proposed that CO molecules are preferentially adsorbed at the hydrophobic region over the surface electrode while the OH molecules can only be adsorbed in the hydrophilic region. Hence, the CO and OHads diffusions are slower, and the coverage within the hydrophobic CO islands remains essentially unchanged. The presence of the membrane adds an additional restriction to the CO oxidation. Although it has been shown that a lower packing density in the CO adlayer favors the CO oxidation,38 in the present investigation conditions this factor apparently plays only a secondary role. It seems that both stronger dipolar coupling and a lower pH near to the surface primarily determine the slower catalytic activity at the Pt(111)/Nafion interface. Conclusions In this work, some structural aspects of the Nafion/Pt(111) interface and its electrochemical behavior have been studied by cyclic voltammetry and in situ infrared reflection absorption spectroscopy (IRRAS). Nafion film showed good stability and electrode adherence, while its electrochemical behavior was modified quantitatively. The membrane introduces a new charge transfer process at 0.50 V and a surface blockage at both hydrogen and OH adsorption regions, due to the membrane/ electrode interactions, strongly dependent on solution pH and the pretreatment methods used in the cast film preparation. IR analysis demonstrates that membrane functional groups respond with the changes in the electrode potential, especially the water and sulfonic entities inside the membrane. In addition, the electric field can drive the film morphology to quasi-irreversible metastable states with enhanced proton transport. The interaction between sulfonic groups of the Nafion membrane and a Pt(111) electrode resembles the interaction between this electrode and a sulfuric acid solution. Additionally, CO stripping experiments reveal that Nafion film modifies the CO-adlayer structure, which only shows on-top adsorption with a higher wavenumber band that continuously increases with the potential, even during CO oxidation, which is slowed in the presence of the ionomer. Finally, the CO oxidation peak for the stripping reaction is sharper and shifted to higher potential values on covered Pt(111) electrode. Acknowledgment. This work has been financially supported by the Generalitat Valenciana (Feder) through project PROMETEO/2009/45. The research of A.M.G.M. has been made possible by a fellowship of the National University of Colombia and COLCIENCIAS, inside the National Program of Formation in Innovation Leaders (Contract No. 472 of 2007). References and Notes (1) Ayato, Y.; Kunimatsu, K.; Osawa, M.; Okada, T. J. Electrochem. Soc. 2006, 153, A203–A209. (2) Zhdanov, V. P.; Kasemo, B. Electrochem. Commun. 2006, 8, 561– 564.

J. Phys. Chem. C, Vol. 114, No. 47, 2010 20139 (3) Wescott, J. T.; Qi, Y.; Subramanian, L.; Capehart, T. W. J. Chem. Phys. 2006, 124, 134702–134706. (4) Malevich, D.; Zamlynny, V.; Sun, S. G.; Lipkowski, J. Z. Phys. Chem. 2003, 217, 513–525. (5) Broka, K.; Ekdunge, P. J. Appl. Electrochem. 1997, 27, 117–123. (6) Buzzoni, R.; Bordiga, S.; Ricchiardi, G.; Spoto, G.; Zecchina, A. J. Phys. Chem. 1995, 99, 11937–11951. (7) Heitner-Wirguin, C. J. Membr. Sci. 1996, 120, 1–33. (8) Ludvigsson, M.; Lindgren, J.; Tegenfeldt, J. Electrochim. Acta 2000, 45, 2267–2271. (9) Gruger, A.; Re´gis, A.; Schmatko, T.; Colomban, P. Vib. Spectrosc. 2001, 26, 215–225. (10) Liang, Z.; Chen, W.; Liu, J.; Wang, S.; Zhou, Z.; Li, W.; Sun, G.; Xin, Q. J. Membr. Sci. 2004, 233, 39–44. (11) Basnayake, R.; Peterson, G. R.; Casadonte, D. J.; Korzeniewski, C. J. Phys. Chem. B 2006, 110, 23938–23943. (12) Basnayake, R.; Wever, W.; Korzeniewski, C. Electrochim. Acta 2007, 53, 1259–1264. (13) Korzeniewski, C.; Adams, E.; Liu, D. Appl. Spectrosc. 2008, 62, 634–639. (14) Byun, C. K.; Sharif, I.; DesMarteau, D. D.; Creager, S. E.; Korzeniewski, C. J. Phys. Chem. B 2009, 113, 6299–6304. (15) Idupulapati, N.; Devanathan, R.; Dupuis, M. J. Phys. Chem. A 2010, 114, 6904–6912. (16) Webber, M.; Dimakis, N.; Kumari, D.; Fuccillo, M.; Smotkin, E. S. Macromolecules 2010, 43, 5500–5502. (17) Warren, D. S.; McQuillan, A. J. J. Phys. Chem. B 2008, 112, 10535– 10543. (18) Maruyama, J.; Inaba, M.; Katakura, K.; Ogumi, Z.; Takehara, Z. J. Electroanal. Chem. 1998, 447, 201–209. (19) Jiang, J.; Kucernak, A. J. Electroanal. Chem. 2004, 567, 123–137. (20) Jiang, J.; Kucernak, A. J. Electroanal. Chem. 2005, 576, 223–236. (21) Chu, Y.; Shul, Y. G.; Choi, W. C.; Woo, S. I.; Han, H. J. Power Sources 2003, 118, 334–341. (22) Tu, W.-Y.; Liu, W.-J.; Cha, C.-S.; Wu, B.-L. Electrochim. Acta 1998, 43, 3731–3739. (23) Miyatake, K.; Omata, T.; Tryk, D. A.; Uchida, H.; Watanabe, M. J. Phys. Chem. C 2009, 113, 7772–7778. (24) Lawson, D. R.; Whiteley, L. D.; Martin, C. R.; Szentirmay, M. N.; Song, J. I. J. Electrochem. Soc. 1988, 135, 2247–2250. (25) Gottesfeld, S.; Raistrick, I. D.; Srinivasan, S. J. Electrochem. Soc. 1987, 134, 1455–1459. (26) Floriano, J. B.; Ticianelli, E. A.; Gonzalez, E. R. J. Electroanal. Chem. 1994, 367, 157–160. (27) Ayad, A.; Naimi, Y.; Bouet, J.; Fauvarque, J. F. J. Power Sources 2004, 130, 50–55. (28) Zecevic, S. K.; Wainright, J. S.; Litt, M. H.; Gojkovic, S. J.; Savinell, R. F. J. Electrochem. Soc. 1997, 144, 2973–2982. (29) Yano, H.; Higuchi, E.; Uchida, H.; Watanabe, M. J. Phys. Chem. B 2006, 110, 16544–16549. (30) Ohma, A.; Fushinobu, K.; Okazaki, K. Electrochim. Acta 2010, 55, 8829-8838. (31) Osawa, M.; Nakane, T.; Ito, K.; Suetaka, W. J. Electroanal. Chem. 1989, 270, 459–464. (32) Kanamura, K.; Morikawa, H.; Umegaki, T. J. Electrochem. Soc. 2003, 150, A193–A195. (33) Malevich, D.; Li, J.; Chung, M. K.; McLaughlin, C.; Schlaf, M.; Lipkowski, J. J. Solid State Electrochem. 2005, 9, 267–276. (34) Subbaraman, R.; Strmcnik, D.; Stamenkovic, V.; Markovic, N. J. Phys. Chem. C 2010, 114, 8414–8419. (35) Subbaraman, R.; Strmcnik, D.; Paulikas, A. P.; Stamenkovic, V.; Markovic, N. ChemPhysChem 2010, 11, 2825–2833. (36) Iwasita, T.; Nart, F. C. Prog. Surf. Sci. 1997, 55, 271–340. (37) Iwasita, T.; Rodes, A.; Pastor, E. J. Electroanal. Chem. 1995, 383, 181–189. (38) Orts, J. M.; Louis, E.; Sander, L. M.; Feliu, J. M.; Aldaz, A.; Clavilier, J. Surf. Sci. 1998, 416, 371–383. (39) Clavilier, J.; Armand, D.; Sun, S. G.; Petit, M. J. Electroanal. Chem. 1986, 205, 267–272. (40) Berna´, A.; Kuzume, A.; Herrero, E.; Feliu, J. M. Surf. Sci. 2008, 602, 84–89. (41) Berna´, A.; Feliu, J. M.; Gancs, L.; Mukerjee, S. Electrochem. Commun. 2008, 10, 1695–1699. (42) Berna´, A.; Climent, V.; Feliu, J. M. Electrochem. Commun. 2007, 9, 2789–2794. (43) Naegeli, R.; Redepenning, J.; Anson, F. C. J. Phys. Chem. 1986, 90, 6227–6232. (44) Elliott, C. M.; Redepenning, G. G. J. Electroanal. Chem. 1984, 181, 137–152. (45) Ferry, L. J. Macromol. Sci., Part A: Pure Appl. Chem. 1990, 27, 1095–1107. (46) Allahyarov, E.; Taylor, P. L. Phys. ReV. E 2009, 80, 020801.

20140

J. Phys. Chem. C, Vol. 114, No. 47, 2010

(47) Allahyarov, E.; Taylor, P. L.; Lo¨wen, H. Phys. ReV. E 2010, 81, 031805. (48) Ma, C.; Zhang, L.; Mukerjee, S.; Ofer, D.; Nair, B. J. Membr. Sci. 2003, 219, 123–136. (49) Kreuer, K. D. J. Membr. Sci. 2001, 185, 29–39. (50) Spry, D. B.; Fayer, M. D. J. Phys. Chem. B 2009, 113, 10210– 10221. (51) Hensley, J. E.; Way, J. D.; Dec, S. F.; Abney, K. D. J. Membr. Sci. 2007, 298, 190–201. (52) Wang, Z.-T.; Wang, Z.-S.; Xu, L.; Gao, Q.-J.; Wei, G.-Q.; Lu, J. J. Power Sources 2009, 186, 293–298. (53) Lin, H.-L.; Yu, T. L.; Han, F.-H. J. Polym. Res. 2006, 13, 379– 385. (54) Osawa, M.; Tsushima, M.; Mogami, H.; Samjeske, G.; Yamakata, A. J. Phys. Chem. C 2008, 112, 4248–4256. (55) Sethuraman, V. A.; Weidner, J. W.; Haug, A. T.; Motupally, S.; Protsailo, L. V. J. Electrochem. Soc. 2008, 155, B50–B57. (56) Lachenwitzer, A.; Li, N.; Lipkowski, J. J. Electroanal. Chem. 2002, 532, 85–98. (57) Chang, S. C.; Weaver, M. J. J. Chem. Phys. 1990, 92, 4582–4587.

Gómez-Marín et al. (58) Lebedeva, N. P.; Koper, M. T. M.; Herrero, E.; Feliu, J. M.; Van Santen, R. A. J. Electroanal. Chem. 2000, 487, 37–44. (59) Furuya, N.; Motoo, S.; Kunimatsu, K. J. Electroanal. Chem. 1988, 239, 347–352. (60) Kitamura, F.; Takeda, M.; Takahashi, M.; Ito, M. Chem. Phys. Lett. 1987, 142, 318–321. (61) Rodes, A.; Pastor, E.; Iwasita, T. J. Electroanal. Chem. 1994, 369, 183–188. (62) Rodes, A.; Pastor, E.; Iwasita, T. J. Electroanal. Chem. 1994, 373, 167–172. (63) Rodes, A.; Pastor, E.; Iwasita, T. J. Electroanal. Chem. 1994, 377, 215–220. (64) Tornquist, W.; Guillaume, F.; Griffin, G. L. Langmuir 1987, 3, 477–482. (65) Hung, L. H.; Wieckowski, A.; Weaver, M. J. J. Phys. Chem. 1988, 92, 6985–6990. (66) Wagner, F. T.; Moylan, T. E.; Schmieg, S. J. Surf. Sci. 1988, 95, 403–428. (67) Kucernak, A. R.; Offer, G. J. Phys. Chem. Chem. Phys. 2008, 10, 3699–3711.

JP107641R